You are on page 1of 22

Chapter 1

INTRODUCTION
The focus of this text is interfaces and colloids, and it makes the
argument that this is the bridge to nanoscience. It often appears that finite
systems, even at equilibrium, cannot be described neatly in terms the usual
set of intensive and extensive variables, i.e., size matters. The most basic
reason for this is that all tangible material systems have boundaries, or
interfaces, that have properties of their own. Physical, as opposed to
mathematical, interfaces are nano-scale strata of material whose structure is
profoundly different from that of bulk phase material, and as the objects of
interest are made smaller and smaller, more and more of their total mass is
part of the interface, with major consequences for their properties. Also, as
the kinetic units are made smaller, their thermal energy assignments begin
to exert influence, and ultimately one must acknowledge the corpuscular
nature of matter: molecules, atoms, sub-atomic particles, etc. How do these
regions blend into one another? It is the in-between region that is of
concern in this text, and it too finds itself divided into parts. In the first,
continuum concepts are retained, but a range of new properties and
phenomena emerge, such as the influence of droplet size on vapor pressure,
spontaneous clumping together of particles, Brownian motion, etc. This is
the traditional domain of colloid science, and extrapolations from it into
the nano region (down to as small as 1-10 nm) often provide valid
descriptions of properties and behavior. Such extrapolations, dealt with in
this text, constitute generic nanoscience. But one may also encounter other
behavior, if continuum descriptions start to break down. Most notable are the
properties attributable to quantum confinement, and it is sometimes to these
kinds of systems that the term nanoscience is restricted. Despite their more
exotic attributes, however, the objects in question are still subject to most, if
not all, of the rulebook of generic nanoscience, the bridge to which is
knowledge of interfaces and colloids. The following brief overview seeks to
set the stage.

A. Interfaces
Interfaces (or surfaces) are the thin boundary regions separating
macroscopic chunks of matter from their surroundings or from one another.
Interface is the more general term for any phase boundary, while the word
surface generally refers to the boundary between a condensed phase and a
gas. Use of the term interface for all situations would be desirable because
it reminds us that the properties of all boundaries depend on both phases.

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTERFACES & COLLOIDS

The surface of a silica micro-particle in contact with air, for example, is


different from its surface in contact with water, or in contact with an
apolar solvent or in vacuo. While interfaces appear abrupt to the naked eye,
they all have in fact a finite thickness, typically ranging from a few
ngstrms to a few nanometers, as suggested in Fig. 1-1. The material in

Fig. 1-1: Schematic of material interface,


showing zone of inhomogeneity and
adjacent bulk phases.

these thin regions is first of all inhomogeneous, and its properties differ
profoundly from those of material in a bulk phase state. No system is free of
the influence of its interface(s), and in many applications or situations, it is
essential that one know not just the bulk properties, but the interfacial
properties as well. For example:

Surface tension determines the shape of small fluid masses or


menisci.

Surfaces of adjacent systems determine whether they stick to, slide


over, or repel one another.

It is often only the interaction of light with the surfaces of systems


that we see.

It is the dependence of surface tension on system composition that


makes possible the formation of such unlikely liquid structures as
foams or froths.

It is often the accumulation of trace components at surfaces, i.e.,


adsorption, that governs the chemistry of their interaction with
adjacent phases.

Consider some of the most fundamental ways in which material in


interfaces differs from bulk materials. First, the state of internal mechanical
stress in the interfacial stratum is different from that in the bulk phases. In
the case of fluid interfaces, this difference is manifest as a measurable
contractile tendency, viz., the surface or interfacial tension. A fluid interface
acts as an elastic membrane seeking a configuration of minimum area. The
origin of the boundary tension can be stated in terms of local internal
mechanical stresses as follows. In the interior of unstrained bulk fluid at rest,
the stress at any point is given by a single scalar quantity, viz., the pressure
p. But in the fluid interfacial layer under the same conditions, the state of
stress requires a tensor for its specification; the components pertaining to

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

stresses parallel to the layer are equal to one another but less than the
component normal to the layer. It is this tangential pressure deficit in the
interfacial layer that leads to boundary tension. The study of fluid interfaces
is referred to as capillarity. Solidfluid and solid-solid interfaces are also
generally in a state of tension, although it is not readily measurable. These
boundary strata have more complicated stress fields than fluid interfaces,
and often are not in states of internal equilibrium. Thus they depend not just
on the thermodynamic state but also on how the interface was formed,
whether from the fracture of a bulk phase, from the stretching of a preexisting surface, from precipitation out of solution, or from some other
process.
The chemical composition of the interfacial stratum is also different
from that of the bulk phases it separates. Consider, for example, the apparent
interface between a piece of metal, such as iron or aluminum, and air, as
shown in Fig. 1-2. Unless it is a noble metal, it will be covered with an oxide
layer, which is at least partially hydroxylated from contact with water vapor
in the air. The -OH groups have an acid or a base character depending on the
metal. On top of this will probably be a layer of tightly bound water and
possibly an additional layer of loosely bound water, depending on the
relative humidity. The very top layer is likely to be an adsorbed scum of
grease or other organic contaminant. The ability of interfaces of all kinds to
contain components that may or may not be present in the bulk phases at the
instant of interest makes their composition more complex than might be
expected.

Fig. 1-2: Schematic representation of the typical chemistry at a nominal


metal-air interface.

The interfacial stratum also often exhibits electrical charge


separation. Interfaces dividing electrically neutral bulk phases may appear
to bear a charge when viewed from one side or the other because positive
and negative charges separate in the direction normal to the interface,
leading to the formation of an electrical double layer. There are a number
of mechanisms by which the charge separation arises, and its existence has
many consequences. Particles dispersed in water, for example as shown in

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTERFACES & COLLOIDS

Fig. 1-3, may repel one another upon close approach, due to overlap of the
like-charged outer portions of the double layers at their surfaces, keeping the
particles from clumping together, and when placed in an external electric
field, migrate by the process of electrophoresis. Fluid interfaces with
electrical charge separation in an external electric field may become unstable
to wavy disturbances, such that they are distorted or even torn apart in the
processes of electro-spraying or electro-emulsification.

Fig. 1-3: Electrical charge separation at the interfaces of particles


against their dispersion medium, water.

The unique mechanical, chemical and electrical properties of


interfaces exert great and highly varied effects upon the behavior of material
systems that cannot be described or explained in terms of bulk phase
behavior alone. Interfacial effects reserve their greatest impact for systems
with large area-to-volume ratios, such as thin films, fine fibers and small
particles, or pushed to the limit: nanofilms, nanowires or nanorods
and nanoparticles.

B. Colloids
Colloids refer to dispersions of small particles, usually with linear
dimensions from 1 nanometer to 10 micrometers, thus spanning the nano
to micro size range. The particles may be either dissolved
macromolecules or macromolecular structures formed from smaller
structural units, or they may constitute a separate phase, as in aerosols,
powders, pigment dispersions, emulsions, micro-foams and finely pigmented
plastics. The description of multiphase colloids, such as those just named,
must take account of the properties of both phases as well as the interface
between them, so that their investigation is a natural adjunct to the study of
interfaces. Reaching down to the size of colloid particles, there are six overarching aspects that distinguish their behavior from that of their larger
counterparts:

Mobility due to thermal kinetic energy


Absence of inertial effects
Negligibility of gravitational effects
Inter-system molecular interactions: adhesion
Size effects on thermodynamic properties, and
Interaction with electromagnetic radiation


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

Simple kinetic theory assigns to all objects (kinetic units), whether


they be molecules, bowling balls or planets, a fixed amount of translational
kinetic energy, viz. 3/2 kT on the average, where k is Boltzmanns constant,
and T is the absolute temperature. This amounts at room temperature to
about 10-20 Joules, completely negligible for objects of macroscopic size. For
molecules in a gas phase, however, it is sufficient to cause them to fly about
in straight lines at velocities of several hundred m/s, colliding with each
other and their confining boundaries. In an ideal gas at room temperature
and atmospheric pressure, the average distance between collisions, the
mean free path, ranges from 10-100 nm, depending on molecular size. In
liquids and solids, the mean free path is of the order of one ngstrm
(0.1nm), and the motion of molecules resulting from their intrinsic kinetic
energy resembles more closely vibration within a cage of nearest neighbors,
with an occasional escape. For colloid particles, usually consisting of
hundreds up to 1010 molecules or so, kinetic effects are still important. They
are sufficiently small that when they are dispersed in a gas or liquid medium,
any unevenness in the bombardment they receive from the surrounding
molecules causes them to move about in a process is known as Brownian
motion. Thus in contrast to macroscopic objects, colloid particles do not
stay put, and the smaller the particles, the more pronounced is their
Wanderlust. The effect of kinetic energy on other objects of colloidal
dimensions, such as ultra-thin deformable films (e.g., bilayer lipid
membranes) or long nanowires, is also evident in the wave-like undulations
these objects exhibit. The effect of intrinsic kinetic energy on fluid interfaces
is manifest as microscopic waves called riplons.
The second consequence of smallness is the absence or near absence
of inertial effects. The response of colloids to the presence of external fields
(gravitational, electric, magnetic, etc.) that act on the particles to orient
them, move them about and generally to concentrate them in some region of
the system is effectively instantaneous. This derives from the law that
Impulse = Momentum, i.e., Ft = mv, where F is the average force acting on
the object during the time interval t, m is the mass of the object, and v is
velocity it attains. Since the mass of a colloid particle is so small, the time
required for it to reach its steady state velocity upon application of a force is
essentially nil. For example, a one-micrometer diameter sphere of density
2.0 g/cm3 sedimenting in water reaches its terminal velocity in about 50
microseconds. The movements of small particles resulting from of an
external field are termed phoretic processes, and may be exemplified by
electrophoresis of particles with electrical double layers in an electric field,
or magnetophoresis of magnetic particles in a magnetic field. The ability to
herd colloid particles around through the control of external fields
provides one of the important strategies for controlling colloidal systems. As
the particles get down into the nano size range, the randomizing effects of
Brownian motion may start to overtake the phoretic processes, and one must
be able to deal with or perhaps even exploit the balance between the two.


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTERFACES & COLLOIDS

The third difference between ultra-small objects and their larger


counterparts is the relative unimportance of gravity compared with other
forces acting on them. Gravitational forces scale with the third power of the
objects linear dimensions, while many of the colloid forces scale with the
second power of the linear dimensions, i.e., they are surface forces.
The fourth difference is the effect that intermolecular interactions
have upon them. Just as molecules interact across a distance with one
another through a variety of forces (dispersion, dipolar, electrostatic, etc.),
objects consisting of many molecules interact with one another by virtue of
the collective effect of these molecular interactions. Particles feel each
other across distances up to tens of nms or greater. Since this is much
greater than the range of interaction between molecules, they are termed
long range forces. The interaction between like particles in a given
medium is always attractive, so that they are drawn together and stick, and a
significant part of colloid science is the design and use of strategies to
prevent such sticking. These rely on repulsive interactions, which are often
the result of electrical charges or the presence of adsorbed macromolecules
at the surfaces. The intervening fluid between the particle surfaces may be
viewed as a film which, upon reaching a certain degree of thinness, may
spontaneously thin further, re-thicken or break up into drops or bubbles.
Such behavior is critical to the formation of ultra-thin coatings as well as to
the drawing together (aggregation and coalescence) of colloid particles in
fluid media.
The fifth distinguishing feature is that many intensive thermodynamic
properties begin to change with particle size as size is reduced. For example,
the vapor pressure of a tiny droplet of liquid depends on its diameter. The
vapor pressure of a 1-m droplet of water at room temperature is
approximately 1% higher than the handbook value. For 100-nm droplets, it
is 12% higher, while for 10-nm droplets it is higher by a factor of three. The
vapor pressure enhancement is one manifestation of the effect of curvature
on thermodynamic properties, known as the Kelvin effect. Other
consequences of the Kelvin effect include the condensation of vapors into
small pores, crevices and capillaries at partial pressures below their vapor
pressure, the enhanced solubility of small particles over larger ones, and the
mechanism of phase change by nucleation. The Kelvin effect governs the
process of cloud formation and ultimately their evolution into raindrops.
The sixth distinction concerns the interaction of electromagnetic
radiation, in particular visible light, with small objects, an interaction
dependent in part on the ratio of the particle size to the wavelength. For
objects large relative to the wavelength, light is refracted or absorbed by the
system, or reflected from its surface and diffracted at its edges, leading to a
blurring of the edges by an amount approximating the wavelength. Thus
ordinary optical microscopy produces images with a resolution no better
than that wavelength (400700 nm). Particles in the colloidal domain scatter


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

visible light, giving a colloid a turbid appearance. Turbidity is a maximum


when the scattering centers (particles) are roughly comparable in size to the
wavelength of visible light, while for smaller particles, falls off as the sixth
power of the particle diameter. For diameters less than approximately 50 nm,
they may become effectively invisible. Most sunscreens use particles that
absorb ultraviolet light but are in a size range that only negligibly scatters
visible light and are thus clear on the skin. In another example, nanoparticles
may be incorporated into optical coatings to improve mechanical properties
(scratch resistance, etc.) without compromising transparency.
It is fun to think of questions regarding the behavior of every-day
systems or processes that cannot be answered without knowledge of
interfaces and colloids. A few examples are listed in Table 1-1. They are
posed here without answers, but all are dealt with in the text. The first
question concerns the fact that one may easily float a needle or other dense
object, such as a paper clip, on water if it gently placed on the water surface.
Table 1-1: Examples of questions that can be answered only with
knowledge of interfacial and colloid science.
1. How can a metal needle (7-8 times as dense as water) be made to float on
water?
2. Why will a teaspoon of certain materials spread spontaneously over several
acres of water surface, and then suppress both waves and evaporation?
3. Why do liquids stick to some surfaces but not to others? How does an
adhesive work?
4. How do soaps and detergents help us to wash things?
5. How can water remain as a liquid at temperatures more than 20C below its
freezing point?
6. How can we make water into a froth or a foam?
7. How can we dissolve large amounts of oil in water using just a trace of a third
component?
8. How can particles much denser than water be suspended in water almost
indefinitely?
9. How can pumping gasoline through a hose lead to a spark, and possible
disaster?
10. How does an absorbent paper towel soak up spills?
11. Why does a liquid jet break up into droplets?
12. How does the addition of salt to turbid water cause it to clarify?

It is evident that factors other than gravity are involved. The second question
is illustrated in the frontispiece of the pioneering monograph by Davies and
Rideal: Interfacial Phenomena, and is reproduced in Fig. 1-4. It shows the
mirror-like surface of Loch Laggan, in Scotland, resulting from the
application of a small amount of hexadecanol, which spreads out into a

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTERFACES & COLLOIDS

monolayer and damps all small wavelets. The monolayer also significantly
suppresses evaporation. Question 3 addresses the everyday observation that
liquids appear to stick to some surfaces, but not others. Teflon-coated
cookware is designed to avoid the sticking of food; carpet fibers are surface
treated with anti-soil coatings; Gore-TexTM outer garments are designed to
shed water. On the other hand, liquids may be formulated to be pressuresensitive adhesives. Question 4 notes the everyday need for soaps and
detergents to facilitate the cleaning of clothes, dishes and our hands, and
asks: What is it that makes the dirt come loose and disappear? Question 5
reminds us that the freezing point of water, or any liquid, found in the
Handbook, does not necessarily tell us the temperature at which freezing

Fig. 1-4: A monolayer of hexadecanol spread on Loch Laggan, Scotland. From: [Davies,
J.T., and Rideal, E.K., Interfacial Phenomena, Academic Press, New York, 1961.]

will actually be observed. If one is careful, for example, water may be


chilled to nearly -40C without freezing, but if subjected to even a small
disturbance, sudden phase change occurs. Question 6 asks why water can
form itself into the delicate polyhedral structure known as a froth or foam,
but apparently only if the water is dirty. Question 7 addresses the experiment
in which water containing dissolved detergent is then capable of dissolving
significant amounts of an otherwise-insoluble oil, such as gasoline. Question
8 is exemplified by an unsettled dispersion of gold particles [sp. gr. 13.7] in
water prepared in Michael Faradays lab in the mid 1800s, as shown in Fig.
1-5, and which can be viewed in the British Museum in London today.
Question 9 recalls a problem associated with the pumping of volatile fuels,
such as during the gassing of an automobile. If the hose is not grounded, or
the gasoline does not have additives giving it a certain degree of electrical


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

conductivity, a potential may develop along the hose great enough to


produce a spark, which in the fuel-air mixture can lead to an explosion.
Question 10 notes the development of absorbent products capable of
imbibing many times their weight in water, and asks how this can be
achieved. Question 11 addresses the fact that cylinders of liquid are unstable

Fig. 1-5: Gold sol prepared by


Michael Faraday, on view in the
Faraday Museum, London.

and will break up into droplets, which in jets occurs too rapidly to be seen by
the naked eye. Figure 1-6 shows a flash photograph of the phenomenon.
The clarification of a turbid dispersion of clay or silt in water by adding salt,
as suggested in Question 12, is observed on a large scale in the formation of
river deltas bordering on saltwater bodies. Looking at an atlas of maps will
confirm that similar deltas do not form where rivers empty into freshwater
bodies.

Fig. 1-6: Breakup of a capillary jet.


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

10

INTERFACES & COLLOIDS

C. The bridge to nanoscience


1. What is nanoscience?
A search of amazon.com for books, using the keyword nanoscience,
reveals 2,061 (!) titles currently (Summer, 2009) in print. The first dozen are
listed in Table 1-2. Even a cursory perusal of the literature reveals the
Table 1-2: The first 12 titles listed by amazon.com with a search request of
nanoscience.)

Introduction to Nanoscience by Gabor L. Hornyak, H.F. Tibbals, Joydeep Dutta,


and Anil Rao (2008)

Nanophysics and Nanotechnology: An Introduction to Modern Concepts in


Nanoscience by Edward L. Wolf (2006)

Nanoscience: Nanotechnologies and Nanophysics by Claire Dupas, Philippe


Houdy, and Marcel Lahmani (2006)

An Introduction to Nanosciences and Nanotechnology by Alain Nouailhat


(2008)

Introduction to Nanoscience and Nanotechnology by Gabor L. Hornyak, H.F.


Tibbals, and Joydeep Dutta (2008)

Introduction to Nanoscale Science and Technology (Nanostructure Science


and Technology) by Massimiliano Di Ventra, Stephane Evoy, and James R.
Heflin (2004)

Nanotechnology For Dummies by Richard D. Booker and Mr. Earl Boysen


(2005)

Nanotechnology: A Gentle Introduction to the Next Big Idea by Mark A.


Ratner and Daniel Ratner (2002)

Understanding Nanotechnology by Scientific American and editors at


Scientific American (2002)

Handbook of Nanoscience, Engineering, and Technology, Second Edition by


William A. Goddard III, Donald W. Brenner, Sergey Edward Lyshevski, and
Gerald J. Iafrate (2007)

Nanotechnology: Basic Science and Emerging Technologies by Mick Wilson,


Kamali Kannangara, Geoff Smith, and Michelle Simmons (2002)

Nanosciences: The Invisible Revolution by Christian Joachim and Laurence


Plevert (2009)

amazing diversity and inter-disciplinary character of this emerging field,


making it somewhat difficult to visualize it as a single, coherent body of
knowledge. In its simplest conception, nanoscience may be defined as the
study of objects with structural elements in the size range of 1100 nm, i.e.,
the low end of the colloidal domain. As indicated earlier, many of the sizedependent aspects of the behavior of entities in the colloidal size range may
be successfully extrapolated into the nano range. The point at which
continuum concepts start to break down is different for different materials
and conditions, and different properties. The bulk-phase equations of state

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

11

may become inapplicable, and the phenomenological equations of transport


(Fouriers Law of heat conduction, Ficks Law of species transport,
Newtons Law of viscosity, Ohms Law, etc.) begin to fail. Thus what may
be required is not just new size-dependent properties, but new physics.
Even though systems requiring new physics are not the focus of this
text, a few examples are recounted here. One of the best known of these is
the observation that the color of nanoparticles of various semiconductors
(called quantum dots) is a function of their size. The electrons in a
quantum dot are confined to distinguishably discrete energy levels (quantum
confinement), dependent on dot size, so that when the dot is illuminated, it
emits only light of a given wavelength, dependent on the particle size and
composition. Figure 1-7 shows a set of dispersions of CdSe/ZnS core-shell
nanocrystals varying in size from  1.7 nm to  5 nm from left to right.

Fig. 1-7: Size-dependent change of the emission color of colloidal dispersions of of


CdSe/ZnS core-shell nanocrystals, varying in size from left to right:  1.7 to  5 nm. For
full color, see image at: [http://www.nanopicoftheday.org/2003Pics/QDRainbow.htm]

This property gives rise to many potential applications, including use as


biological markers. Quantum dots of a known range of size and
concentration might be incorporated into a single colloidal latex particle,
which in turn may be attached to a particular type of tissue using an
appropriate biological targeting agent. Upon illumination, such a particle
will produce a unique spectrum (a spectral bar code) characteristic of the
particular collection of quantum dots, thus identifying the existence and
location of the tissue in question. As another example, the transport of
electricity along a nanowire or a carbon nanotube (virtually the poster child
for nanotechnology) is a function of diameter and may be governed by
quantum effects. This means it may be possible to construct transistors,
diodes, switches, gates, as well as conductors and other components of
microcircuits from such objects, leading to ultra-small electronic computer
logic systems. Recent discussion has moved to the possibility of using
properly designed single molecules for this purpose. One of the problems of
using these tiny elements for controlling charge flow is the fact that

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

12

INTERFACES & COLLOIDS

whenever they are bumped by neighboring nano objects (kinetic energy!),


they are easily discharged.
Another example of the specific importance of size in the nano range
concerns magnetic properties. A particle of magnetite (Fe3O4) in this range,
for example, will retain its magnetism more effectively the larger it is,
making it a better magnetic data storage medium than smaller particles,
which are subject to loss of their magnetic information by being bumped by
their neighbors. When it becomes too large, however, it will split into two
magnetic domains of opposite polarity, and much reduced total magnetism.
Thus there is an optimum single-domain magnetic nano-crystal size (called
the super-paramagnetic limit, and often about 50 nm) for data storage, well
known to the manufacturers of magnetic data storage media. Other examples
of the special photonic, electronic and magnetic properties of nano-objects
could be listed.
2. Nanostructures and assemblies
In the nano range, one often cannot refer to the objects of interest as
phases or particles, but must think of them as structures. Sometimes
nanoparticles are single molecules, or consist of a countable number of
atoms or molecules, arranged in particular ways. It is useful to describe just
a few of these nano-structures to give an idea of what is possible.
Buckyballs: As pictured in Fig. 1-8, Buckyballs are hollow spherical
molecules of carbon atoms, more formally known as Buckminsterfullerene, in view of their resemblance to structures created by the architect
and inventor, Buckminster Fuller (1895-1983). The structure first discovered
in 1985 upon examining the debris formed by vaporizing graphite with a
laser was C60. It consists of 5-member rings isolated by 6-member rings and
is approximately one nm in diameter. Since then, many other fullerenes of
different numbers of carbon atom and different structures have been
discovered and characterized. Furthermore, fullerene-like structures of other

Fig. 1-8: Two views of the structure of a C60 Buckyball


From: [www.bfi.org/?q=node/351]


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

13

materials, notably tungsten and molybdenum sulfides, WS2 and MoS2, have
recently been made and characterized. Atoms or small molecules can be put
inside these structures or attached to their exteriors. These structures have
some remarkable properties. For example, when certain metal atoms are
trapped inside the cages, they exhibit superconductivity.
Carbon nanotubes: As shown in Fig. 1-9, carbon nanotubes are
cylinders whose walls are made from monomolecular sheets of graphite,
termed graphene, resembling chicken wire. Graphene layers themselves,
including their exfoliation from graphite and their bottom-up synthesis, are
objects of current interest. The tubes may be single-walled (SWNTs), about
1.5 nm in diameter, or multi-walled nanotubes (MWNTs) of somewhat
larger diameter. Different types of SWNTs are formed when the graphene

Fig. 1-9: Carbon-wall nanotubes,


showing single-walled (SWNT)
and multi-walled (MWNT)
configurations an a tube filled with
nanoparticles. From:
[www.wtec.org/loyola/nano/04_03
.htm]

sheet is rolled in different ways, and the different structures have different
electrical properties. One type is conductive, one is semi-metallic and
another is a semiconductor, with quantum confinement of electrons. They
may be filled with metal or other atoms and chemically functionalized
externally. Carbon nanotubes are usually formed in a plasma arcing process,
although other methods of synthesis are available, and still others are under
investigation. They are incredibly strong for their size and weight, and may
be the basis for super reinforced nanocomposites. It has even been
speculated that they may be the basis for a cable supporting a space
elevator on which one might ride to the stratosphere. Nanotubes can also be
made from other inorganic materials producing structures different from
those of carbon and holding the promise of a range of new applications.
Vesicles: As shown in Fig. 1-10, vesicles are spherical bilayer
structures of lipids or other surfactants, and may be single-walled or multiwalled, with an aqueous interior. They form spontaneously in water upon
gentle agitation when the right monomeric species (usually di-tail surfactants
or lipids) are present. Vesicles may be as small as 30 nm in diameter, or as
large as one micrometer or more. The smaller ones in particular are showing
promise as drug delivery vehicles, with an appropriate drug in the interior


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

14

INTERFACES & COLLOIDS

Fig. 1-10: Single-walled


vesicle formed from di-tail
lipid monomers.

and targeting agents attached to their outer surfaces, or nano-reactors of


various sorts.
Dendrimers: As shown in Fig. 1-11, dendrimers are highly ordered,
regularly branched single globular macromolecules, sometimes called
starburst polymers. They are formed by the successive addition of layers (or
generations) of chemical branches, and have reached sizes up to 100 nm.
The sponginess of their structure has led to their use as impact modifiers in
composite materials. These remarkable molecules can also be designed to
allow the incorporation of desired guest molecules in their interiors (for
applications like drug delivery) or chemical functionality of their exteriors.
Their controllable chemistry and structure has led to their use as synthetic
catalysts themselves, and recently a dendrimer-based method has been
reported for controllably synthesizing 1-nm and subnanometer-sized metal
catalyst particles containing a well-defined number of atoms. It turns out that
3-nm diameter particles show higher catalytic activity for the oxygenreduction reaction (ORR) than either smaller or larger particles.

Fig. 1-11: Bengal rose encapsulation in a dendrimer center. From: [Zeng, F. and
Zimmerman, S. C., Chem. Rev., 97, 1681 (1997).]


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

15

Many more nanostructures are displayed in the delightful website:


Nanopicture of the Day: www.nanopicoftheday.org, from which three
more examples are drawn.
Nanoforests: Figure 1-12 shows a scanning electron micrograph
(SEM) of a forest of uniform gold nanowires produced by plasma etching
a polycarbonate film containing the nanowires. The loaded film had been

Fig. 1-12: A forest of gold


nanowires produced by plasma
etching a polycarbonate film
containing the nanowires.
From: [Yu, S., Li, N., Wharton,
J., and Martin, C. R., Nano
Letters, 3, 815 (2003).]

prepared by electrochemically filling holes that had been etched in the


polycarbonate by electron beam lithography. When the nanowirescontaining membranes are exposed to an O2 plasma, the polymer at the
membrane surface is selectively removed, exposing the ends of the gold
nanowires. The length of the nanowires is controlled by varying the etch
time. Other types of nanoforests can be constructed in other ways. Tree-like
nanostructures (nanotrees) can be formed through the self-assembled
growth of semiconductor nanowires via a vaporliquidsolid growth mode.
This bottom-up method uses initial seeding by catalytic nanoparticles to
form the trunk, followed by the sequential seeding of branching structures.
This controlled seeding method has potential as a generic means to form
complex branching structures, and may also offer opportunities for
applications. One of the methods for synthesizing carbon nanotubes is by
chemical vapor deposition using catalyst seeds that produce carpets of the
nanotubes fibers.
Nanohelices: Figure 1-13 shows a scanning electron micrograph of
nanohelices formed from single-crystal zinc piezo-electric nano-belts
produced by chemical vapor deposition. They are formed in response to the
electrostatic energy introduced by the spontaneous polarization across the
thickness of the nanobelt, owing to the presence of large polar surfaces. The
shape of the nanohelix is determined by minimizing the energy contributed
by electrostatic interaction and elasticity. The piezoelectric and possibly
ferroelectric nanobelts have potential as surface selective catalysts, sensors,
transducers, and micro-electromechanical devices.


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

16

INTERFACES & COLLOIDS

Fig. 1-13: Nanohelices.


From: [Kong, X. Y., and
Wang, Z.L., Nano Letters,
3, 1265 (2003).]

Nanoarrays: Figure 1-14 shows a regular array of platinum catalyst


nanoparticles, 20 nm in diameter, 15 nm high and spaced 100 nm apart, as
viewed by an atomic force microscope. It is found that catalytic activity
often depends critically on the particle size, aspect ratio and spacing, and
these features can now be controlled in the synthesis process. In the case
shown, the structure was created using a template formed by a thin polymer
film coated onto a smooth silicon oxide surface subsequently etched into the
desired pattern by an electron beam. A platinum film was then evaporated
onto the polymer, filling in the holes. Finally, the polymer film was
removed, leaving the pattern shown.

Fig. 1-14: An array of


platinum catalyst
nanoparticles. From: [Gabor
Somorjai: Nano Picture of
the Day, Jan. 24, 2004]

Structures such as those described above and others may in turn be


dispersed in a medium or assembled (or self-assembled) into mesoscopic or
macroscopic superstructures (monoliths) with unique and important
properties. Micro- or nano-composites may be constructed of units of these
dimensions to produce materials of almost any desired properties; for
example, materials with the heat resistance of ceramics and the ductility of
metals, or materials with the strength of steel, but the weight of a polymer,
etc. Sintered arrays may produce nano-filters or membranes, super thermal
insulation materials, or storage media for liquefied gases at modest
pressures. One exciting possibility is that of constructing nano-circuits from
quantum dots, nanowires and other elements, put into a proper array onto
some surface. Another possibility is the creation of a device for bringing
chemicals together in miniscule quantities so that they may react to produce
some recordable result. One may imagine an assembly of nano- or microsized tubes, channels, reservoirs, etc., assembled on some surface for this

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

17

purpose. These are referred to as LOC (lab-on-a-chip) configurations and


may be the basis for chemical sensors more accurate than any known today
and small enough to be injected into the blood stream. The motion of fluids
through such a system is governed by micro- or nano-fluidics, which in
the ultimate may not be able to treat the fluids as continua. Still other
configurations made of nano building blocks might be motors or mechanical
actuators, converting chemical energy into mechanical work or electricity
with nearly 100% efficiency. These structures, known as MEMS or NEMS
(micro- or nano-electromechanical systems) may someday be the
molecular assembly devices for making structured nano-objects. One
concept for the realization of such devices is that of harnessing the random
kinetic energy of all kinetic units by devising ratchets that allow net
movement to occur only in one direction, and another speculated that they
could be the basis for nano-manufacturing (nanobots), perhaps even to
replicate themselves.
As their potential has come into clearer view over the past decade,
nanoscience and its application, nanotechnology, have become among
the hottest topics for academic inquiry, government research funding and
fanciful speculation. George Whitesides (Professor of Chemistry, Harvard
University), in a recent (March, 2009) plenary lecture before the American
Chemical Society, described nanoscience and nanotechnology as now
entering late adolescence, i.e., past the exaggerated expectations of infancy
(self-replicating nanobots, space elevators, single-molecule computing,
etc.) as well as the over-reaction to the inevitable disappointments, and
moving into a more mature status in which the focus has returned to
fundamentals, discovery, and science.
3. Generic nanoscience
In addition to the dazzling array of special effects associated with
many of the objects of nanoscience, perhaps one of the most amazing
observations is how many of them may be described by just properly taking
into account the generic effects of smallness while extrapolating the world of
interfaces and colloids into the nano domain. Despite any special properties
they might have, all of these micro or nano entities must contend with the six
universal aspects of smallness listed earlier, and it is to these that principal
attention is given in this text. The practical issues concerning these entities
are that they must be:
Produced, either by somehow subdividing larger systems in socalled top-down procedures, or grown out of molecular or
atomic media by so-called bottom-up procedures;
Characterized, i.e., one must be able to see or image the
particles or structures and determine their mechanical,
chemical, electrical, etc., properties


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

18

INTERFACES & COLLOIDS


Assembled, often into two-dimensional or three-dimensional
arrays by means of flow, external fields, or by various selfassembly processes;
Manipulated, i.e., placed where they are wanted, possibly one at
a time, and usually on some surface, in accord with some
blueprint; and
Protected, against moving out of position, becoming attached to
unwanted particles, vaporizing, etc.
4. New tools of generic nanoscience

The recent emergence of nanoscience and nanotechnology owes as


much to the development and refinement of tools for the accomplishment of
the above tasks as it does to the birth of quantum dots, carbon nanotubes,
and the like. Some of these new tools are listed in Table 1-3. One of the
most important developments is the family of scanning probe techniques,
described briefly here but in more detail in Chap. 4.M. The first to emerge
was Scanning Tunneling Microscopy (STM), pictured in Fig. 4-92. When a
sharp conducting tip is poised a few nanometers above a conducting surface,
Table 1-3: Some new tools for the characterization and
manipulation of nanosystems.
SPM (scanning probe microscopy
STM (Scanning Tunneling Microscopy)
AFM (Atomic Force Microscopy)
NSOM (Nearfield Scanning Optical Microscopy)
Colloid force-distance measurement

SFA (Surface Forces Apparatus)


MASIF (Measurement and Analysis of Surface Interactions
and Forces)
TIRM (Total Internal Reflection Microscopy)

Particle manipulation

Optical tweezers (optical trapping)


Traveling-wave dielectrophoresis

and a slight bias is imposed across the gap, electrons may tunnel directly
across. It occurs when the highest occupied molecular orbital (HOMO) of
the material on one side of the gap overlaps with the lowest unoccupied
orbital (LUMO) of the material on the other side of the gap. The tunneling
current is exponentially sensitive to the distance between the tip and the
nearest atom on the surface. As the tip is rastered across the surface, it is
moved up and down by a piezoelectric controller to maintain the current
constant. The up-and-down motion of the controller is tracked and recorded
by a computer to produce a topographical map of the surface with atomic
resolution. A related device, the Atomic Force Microscope (AFM), does not

  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

19

require that the tip and surface be conductive. In the simplest mode of
operation, shown in Fig. 4-93, the sharp tip is held with constant force in
direct contact with the surface by means of an ultra-thin cantilever beam. As
the sample is rastered below the tip, the cantilever moves up and down to
follow the topography. The motion of the cantilever is tracked by means of a
laser reflected from its back to a photodiode system, revealing the
topography of the scanned sample. The topographical resolution of STM or
AFM is of the order of ngstrms, far better than that of the wavelengthlimited optical microscopy. The techniques are thus ideally suited to the
examination of nano features and nano structures. AFM may also be
configured to map adhesion, stiffness, lubricity, and many other properties
of the scanned surface, or may be operated in an intermittent contact mode
to probe soft samples or in a non-contact mode (in which the tip vibrates
above the surface) to probe the force fields emanating from the surface.
Another scanning probe method for overcoming the wavelength
limitations of optical microscopy is that of Nearfield Scanning Optical
Microscopy (NSOM), shown in Fig. 4-91, in which the object of interest is
illuminated by the evanescent light which emerges a few tens of nm from the
back side of a medium in which there is total internal reflection of light.
Information pertaining to a spot a few nm2 is captured by a sharp transparent
probe tip a few nm above the spot and transmitted to a computer. Images are
again built up by rastering the tip over the surface, or the sample beneath the
probe.
A variety of methods have been developed for the direct measurement
of forces of the type existing between proximate colloid particles, as
described in Chap. 7 in detail. One of the most important of these devices is
the Surface Forces Apparatus (SFA), pictured in Fig. 7-1, in which the force
acting between a pair of approaching crossed cylinders (of radii of a few
mm) attached to sensitive leaf springs. Positioning and force measurement
are effected using piezo-electric positioners, and the distance of separation
(with ngstrm resolution) is determined interferometrically. An important
breakthrough in the development of the SFA was the discovery that freshlycleaved Muscovite mica is atomically smooth, and is thus used for the
surfaces of the approaching cylinders. Other materials of interest may be
coated on to these substrates. The MASIF device is similar to the SFA
accept that a bimorph strip (two slices of piezo-electric material sandwiched
together) is used to measure the force between the approaching surfaces.
TIRM (Total Internal Reflection Microscopy) monitors the level of particles
suspended against gravity by repulsive forces above a surface from which an
evanescent beam is emerging.
A number of methods are being developed for the manipulation of
micro- or nano-particles one at a time. The AFM probe may be used to push
or pull nano structures (and even single atoms!) around on a surface. In a
famous image now more than a decade old, IBM researchers in Zrich


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

20

INTERFACES & COLLOIDS

produced the company logo in this way by arranging an array of 35


individual xenon atoms on a Ni(110) surface into the letters IBM. Figure 115 shows AFM images of strands of DNA adsorbed onto a modified
graphite surface. Manipulation was effected by bringing the AFM tip into
contact with the surface and moving it using homemade hardware and
software: (a) shows molecules as deposited, (b) shows two the molecules

Fig. 1-15: Manipulation and overstretching of genes on a solid substrate.


From: [Severin, N., Barner, J., Kalachev, A. A., and Rabe, J. P., Nano
Letters, 4, 577 (2004).]

stretched into diamond shapes, and (c) shows fragments drawn into the word
science. Nano objects can also be picked up using a pair of carbon
nanotubes like chopsticks attached to a cantilever tip. The tweezer-like
device is closed or opened with an applied electrical bias. In a simpler way,
particles can be plucked out of an array, or put down in a desired position on
a surface by charging and discharging a cantilever tip.
Another technique born out of the technology of atomic force
microscopy is Dip-Pen Nanolithography (DPN), a scanning probe nanopatterning technique pictured in Fig. 1-16 in which an AFM tip is used to

Fig. 1-16: Schematic of


dip-pen nanolithography.

deliver molecules to a surface via a solvent meniscus, which naturally forms


in the ambient atmosphere. This direct-write technique offers high-resolution
patterning capabilities for a number of molecular and bio-molecular inks
on a variety of substrate types such as metals, semiconductors, and
monolayer functionalized surfaces. Figure 1-17 shows a paragraph from
Richard Feynmans 1959 speech: Theres Plenty of Room at the Bottom,


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

INTRODUCTION

21
1

written using DPN on a 7-m square surface. Feynmans speech is regarded


as the birth document of nanoscience, and the portion of is shown in Fig. 117 reads:

Fig. 1-17: DPN-written portion


of Feynmans speech: Theres
Plenty of Room at the Bottom.
From: Mirkin Group,
Northwestern University
[http://www.nanotechnow.com/basics.htm]

As soon as I mention this, people tell me about miniaturization, and


how far it has progressed today. They tell me about electric motors that
are the size of the nail on your small finger. And there is a device on the
market, they tell me, by which you can write the Lords Prayer on the
head of a pin. But thats nothing: thats the most primitive, halting step
in the direction I intend to discuss. It is a staggeringly small world that
is below. In the year 2000, when they look back at this age, they will
wonder why it was not until the year 1960 that anybody began seriously
to move in this direction.

A powerful technique for positioning and holding a micro-particle at a


desired position in three-dimensional space is that of optical trapping,
pictured in Fig. 6-44. A particle whose refractive index differs significantly
from that of the suspending medium, and whose size is greater than the
wavelength of the light, is held by photonic forces in the focused beam of a
laser. The device is also referred as laser tweezers, and may be used to
assemble structures one particle at a time or several at a time into aggregates
or arrays of desired structure using independently-controlled, time-shared
optical traps or holographic optical tweezers, in which a propagating laser
wave front is modified by passing through a pattern of interference fringes.
Laser tweezers may also be used as in situ force sensors as well as
micromanipulators. They thus provide, in principle, an additional method for
probing colloid force-distance relationships.
Another method for moving particles is that of traveling wave
dielectrophoresis, in which particles are acted upon by a traveling,
1

Feynman, R. P., Engineering and Science (Caltech), Vol. XXIII, No. 5, pp. 22-36 (1960).


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

22

INTERFACES & COLLOIDS

sinusoidal electric field, which induces dipoles in the particles. When there
is a time lag between the induced dipole and the field, a particle experiences
a force, which induces motion.
5. The plan
The journey in this text begins with a study of fluid interfaces and the
many consequences of the contractile tendency (surface or interfacial
tension) they exhibit, i.e., capillarity. Next is an explicit look at the
thermodynamics of interfacial systems, particularly those that are
multicomponent. Adsorption at both fluid and solid interfaces is examined,
and the wondrous diversity of surfactants and their solutions is investigated.
It moves on to the examination of the physical interaction between liquids
and solids, and the interaction of fluid interfaces with solid surfaces. This
includes wetting, coating, spreading, wicking and adhesion. It includes
examination and characterization of the properties of solid surfaces:
topography, surface energy, and surface chemistry. Next, the world of
colloids, including nano-colloids, is explored. Their morphological, kinetic,
phoretic and optical properties are described. The electrical properties of
interfaces, and their many consequences, are examined next, and the
following chapter deals with the interaction between colloid particles.
Following a consideration of colloid system rheology, fluid phase colloids,
i.e., emulsions and foams, are examined, and the journey concludes with a
consideration of interfacial hydrodynamics, symbolized by the image of
wine tears on the cover.


  

 
  !"#
$!$$
%&''((($(#)$*''+,+-$*

You might also like